ArticlePDF Available

The Dynamic Changes of DNA Methylation and Histone Modifications of Salt Responsive Transcription Factor Genes in Soybean

PLOS
PLOS ONE
Authors:

Abstract and Figures

Epigenetic modification contributes to the regulation of gene expression and plant development under salinity stress. Here we describe the identification of 49 soybean transcription factors by microarray analysis as being inducible by salinity stress. A semi-quantitative RT-PCR-based expression assay confirmed the salinity stress inducibility of 45 of these 49 transcription factors, and showed that ten of them were up-regulated when seedlings were exposed to the demethylation agent 5-aza-2-deoxycytidine. Salinity stress was shown to affect the methylation status of four of these ten transcription factors (one MYB, one b-ZIP and two AP2/DREB family members) using a combination of bisulfite sequencing and DNA methylation-sensitive DNA gel blot analysis. ChIP analysis indicated that the activation of three of the four DNA methylated transcription factors was correlated with an increased level of histone H3K4 trimethylation and H3K9 acetylation, and/or a reduced level of H3K9 demethylation in various parts of the promoter or coding regions. Our results suggest a critical role for some transcription factors' activation/repression by DNA methylation and/or histone modifications in soybean tolerance to salinity stress.
Methylation status of the promoter region of four salinity-responsive TFs in untreated (S0) and salinity-stressed (S1-S24) seedlings (S1: 1h, S3: 3h, S6: 6h, S12: 12h, S24: 24h). (a) The black and white boxes indicate, respectively, exon and untranslated regions. The short bars annotated with ''I, II, III'', ''a'' or b'' indicate, respectively the sequences subjected to ChIP analysis, genomic bisulfite sequencing and those used as probes for Southern blotting. The long vertical bars marked ''c'' display the distribution of CG dinucleotides (marked with red vertical lines), and CNG (blue vertical lines) and CNN (black vertical lines) trinucleotides. The red vertical lines marked with a rectangular indicate CCGG sites analyzed by Southern blotting. The thick black vertical lines represent the proportion of methylated cytosine. Ten positive clones from each gene's amplicon were sequenced. The data reflect the outcome of three independent experiments, and error bars represent standard error (SD). (b) The efficiency of the bisulfite treatment to transform unmethylated cytosine to thymine. A fragment of Glyma20g32730 with numerous cytosines was cloned into Dm-E. coli cells and the plasmid was treated with bisulfite in parallel with the soybean genomic DNA. All clones processed showed a transformation rate .99.7%. (c) Methylation-sensitive DNA gel blot analysis of non-stressed (S0) and salinity-stressed seedlings (S1-S24). Genomic DNA was digested to generate large fragments, then with one or other of the schizomers HpaII or MspI. Hybridization probes indicated. A DNA fragment amplified from the probe sequence was used as a positive control (+), and ddH 2 O was used as a negative control (2). doi:10.1371/journal.pone.0041274.g003
… 
Content may be subject to copyright.
The Dynamic Changes of DNA Methylation and Histone
Modifications of Salt Responsive Transcription Factor
Genes in Soybean
Yuguang Song, Dandan Ji, Shuo Li, Peng Wang, Qiang Li, Fengning Xiang*
The Key Laboratory of Plant Cell Engineering and Germplasm Innovation, School of Life Sciences, Shandong University, Jinan, Shandong, China
Abstract
Epigenetic modification contributes to the regulation of gene expression and plant development under salinity stress. Here
we describe the identification of 49 soybean transcription factors by microarray analysis as being inducible by salinity stress.
A semi-quantitative RT-PCR-based expression assay confirmed the salinity stress inducibility of 45 of these 49 transcription
factors, and showed that ten of them were up-regulated when seedlings were exposed to the demethylation agent 5-aza-2-
deoxycytidine. Salinity stress was shown to affect the methylation status of four of these ten transcription factors (one MYB,
one b-ZIP and two AP2/DREB family members) using a combination of bisulfite sequencing and DNA methylation-sensitive
DNA gel blot analysis. ChIP analysis indicated that the activation of three of the four DNA methylated transcription factors
was correlated with an increased level of histone H3K4 trimethylation and H3K9 acetylation, and/or a reduced level of H3K9
demethylation in various parts of the promoter or coding regions. Our results suggest a critical role for some transcription
factors’ activation/repression by DNA methylation and/or histone modifications in soybean tolerance to salinity stress.
Citation: Song Y, Ji D, Li S, Wang P, Li Q, et al. (2012) The Dynamic Changes of DNA Methylation and Histone Modifications of Salt Responsive Transcription
Factor Genes in Soybean. PLoS ONE 7(7): e41274. doi:10.1371/journal.pone.0041274
Editor: Xiaoyu Zhang, University of Georgia, United States of America
Received May 7, 2012; Accepted June 19, 2012; Published July 18, 2012
Copyright: ß2012 Song et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits
unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Funding: This work was supported by the National Special Science Research Program of China (grant no. 2007CB948203) http://www.973.gov.cn/AreaMana.aspx;
the National Natural Science Foundation of China (grant nos. 30970243 and 30771116) http://www.nsfc.gov.cn; Excellent Youth Foundation of Shandong Province
of China (grant no. JQ200810) http://www.sdnsf.gov.cn/portal/; Science & Technology Plan of Shandong Province (grant 2009GG10002001) http://www.sdstc.gov.
cn/index.jsp and the Chinese Natural Education Ministry Doctor Station Foundation Fellowship (grant no. 913111006) http://std.nankai.edu.cn/kyjh-bsd/1.htm.
The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.
Competing Interests: The authors have declared that no competing interests exist.
* E-mail: xfn0990@sdu.edu.cn
Introduction
Soybean (Glycine max (L). Merr.) is an important source of
protein and oil in both the human and domestic animal diet. As
for most crop species, its productivity is significantly compromised
by soil salinity [1], but, like most plants, it has evolved a variety of
mechanisms to aid its survival under environmental stress. The
expression of many plant genes is altered by salinity stress; some of
these encode aspects of cellular metabolism and stress tolerance,
while others are regulatory in nature [1,2]. Transcription factors
(TFs), which belong to the latter class, have been classified into a
number of families on the basis of their sequence, and some
members of the MYB,NAC,b-ZIP and AP2-DREB families have
been shown to be intimately involved in the stress response
[3,4,5,6,7,8]. Such as, the heterologous expression of three
soybean MYB and three b-ZIP TFs in Arabidopsis thaliana improved
its response to salinity and freezing stress [9,10]. Similarly the
heterologous expression of GmDREB2 was able to enhance the
drought and salinity tolerance of tobacco [11], as did the over-
expression of either GmNAC11 or GmNAC20 for soybean [12].
Once a plant detects the onset of stress, TFs characteristically
respond by inducing the expression of a cascade of downstream
targets. However, their activation is in part also dependent on their
chromatin structure, which is largely determined by epigenetic
means [13,14,15,16]. Cytosine methylation within the promoter
sequence has been shown to underlie numerous cases of gene
down-regulation or silencing [17,18,19,20,21]. DNA methylation
in the plant genome mostly at CG dinucleotides and CNG
trinucleotides, but also at an asymmetrical sequence contexts CNN
(N is any nucleotide but G) [22,23,24]. The N terminus of the
histone molecule can be acetylated, phosphorylated, methylated,
ubiquitinated or ribosylated [25]. The presence of the trimethy-
lated form of histone H3K4 and of the acetylated form of H3K9 in
the promoter region have been frequently associated with
transcriptional activation, while that of the dimethylated form of
H3K9 represses it [26,27,28]. Sometimes, H3K9 methylation can
trigger cytosine methylation in both Neurospora crassa [29] and A.
thaliana [30], while cytosine methylation at the CNG trinucleotide
appears to be partially dependent on the activity of a histone
methyltransferase [29,31,32].
A number of examples where epigenetic modification has
contributed to the regulation of gene expression during periods of
environmental stress have been presented. In particular, the low
temperature induced expression of the maize gene ZmMI1 has
been correlated with a reduction of DNA methylation in its
nucleosome core [33]. In tobacco, several stress agents are known
to promote demethylation in the NtGPDL coding sequence, leading
to alterations in its level of expression [34]. The submergence of
rice seedlings reduces histone H3K4 trimethylation and acetyla-
tion in genes encoding both alcohol dehydrogenase and pyruvate
decarboxylase, leading to their up-regulation [35]. In A. thaliana,
the drought-induced expression of a number of stress-responsive
PLoS ONE | www.plosone.org 1 July 2012 | Volume 7 | Issue 7 | e41274
genes has been associated with an increase in H3K4 trimethyla-
tion and H3K9 acetylation [36]. Here, we set out to document the
induction by salinity stress of DNA methylation and histone
modification in a number of salinity responsive soybean TFs, and
to identify what relationship there is, if any, between the
expression of a TF and the epigenetic status of its promoter
sequence. In addition, since to date no systematic attempt has
been made to investigate the dynamics and reversibility of both
DNA and histone modification over the course of a stress episode,
we have explored this feature focusing on four salinity stress
inducible soybean TFs.
Materials and Methods
Exposure of soybean seedlings to salinity stress and 5-
ADC treatment
Seedlings of the soybean cultivar Williams 82 were grown in
vermiculite under a 16h photoperiod at 25uC for 14 days before
being exposed to stress treatment. Once the seedlings had been
removed from the vermiculite and their roots rinsed in water, they
were then treated with either 150 mM NaCl for 1h, 3h, 6h, 12h or
24h, or with 50 mM 5-aza-2-deoxycytidine (5-ADC) for 12h, 24h,
48h or 72h. RNA and DNA was extracted from snap-frozen plants
both before the stress treatment had begun and then at each time
interval. Mock treatments (ddH
2
O only) were included as a
control. RNA was prepared from 0.2 g plant material using the
TRIzol (Invitrogen) reagent, following the manufacturer’s proto-
col, and DNA was extracted from 1 g plant material using a
DNeasy Plant Mini kit (Qiagen).
Microarray analysis
RNA were isolated from the mock (M0, M1, M3, M6, M12,
M24) and salinity treated (S0, S1, S3, S6, S12, S24) seedlings.
0.5 mg RNA that extracted from each time point of the mock and
salinity-stressed seedlings were mixed respectively to obtain the
mock and salinity-stressed RNA pools, and then they were used to
synthesize the cDNA. The cDNA was labeled with biotin, and
then hybridized to an Affymetrix soybean Genome Array
according to the manufacturer’s instructions (15h in a rotating
hybridization oven set at 45uC and 60 rpm). After the hybridiza-
tion, the microarrays were scanned using a GeneChipHScanner
3000 (Affymetrix, P/N 00-00212). Then the scaling factor,
background, noise, and percentage presence were calculated
according to the Affymetrix Data Mining Tool protocols
(Affymetrix). All resulting datasets were filtered using the absolute
call metric (present or absent) implemented within Microsoft
Access (Microsoft Corporation, Redmond, WA) and the micro-
array data were processed in an R (v2.7.0) environment, using the
LIMMA package [37]. Quantile normalization was performed. A
single repeat microarray analysis for each group was performed.
Figure 1. The expression of the 49 TFs in mock-stressed and salinity-stressed seedlings. (a) GmAP2-DREBs, (b) GmMYBs, (c) GmNACs and
(d) Gmb-ZIPs. M0-M24 refer to seedlings exposed to just ddH
2
O for, respectively, 0h, 1h, 3h, 6h, 12h and 24h; S0–S24 refer seedlings exposed to
150 mM NaCl for 0h, 1h, 3h, 6h, 12h and 24h, respectively. Each gene-specific region was amplified by RT–PCR using the gene-specific primers
(Table S2). The TUBULIN gene (Genbank accession AY907703) was used as an internal control. The experiment was repeated three times with similar
result.
doi:10.1371/journal.pone.0041274.g001
Epigenetic Modifications Affect Salt Response
PLoS ONE | www.plosone.org 2 July 2012 | Volume 7 | Issue 7 | e41274
Transcript level analysis
Semi-quantitative RT-PCR (sqRT-PCR) and quantitative real-
time RT-PCR (qRT-PCR) were employed to quantify transcript
levels more precisely. RNA was extracted from 0.2 g of seedling
material ground in liquid nitrogen by the addition of 1 ml
TRIZOL reagent (Invitrogen) and treated with RNase-free DNase
I. A 3 mg aliquot of total RNA was used to generate the first cDNA
strand with the SuperScript First-Strand Synthesis System
(Invitrogen) according to the manufactuter’s instructions. A 1 ml
aliquot of this cDNA was used as the template for a 22–34 cycle
sqRT-PCR, where the cycling regime was 94uC/30 s, 55uC/30 s,
72uC/30 s. A fragment of the soybean TUBULIN gene (Genbank
accession AY907703) was used as a reference. Primer sequences
are given in Table S2. Each 15 ml qRT-PCR contained 7.5 ml
Maxima SYBR Green qPCR Master mix buffer (Roche), 0.5 ml
10 mM specific primers, 1.5 ml of a 1:10 dilution of cDNA and
5.5 ml ddH
2
O. The cycling regime consisted of a denaturation step
(95uC/3 min) followed by 18–35 cycles of 95uC/30 s, 60uC/15 s,
72uC/15 s, and a fragment of the soybean TUBULIN gene
(GenBank accession AY907703) was used as a internal control.
Primer sequences are given in Table S2. The relative expression
level of the target sequence was determined using the 2
2DDCt
method [38]. Each estimate was derived from the mean of three
independent biological replicates.
Bisulfite DNA sequencing
A2mg DNA aliquot was dissolved in 50 ml ddH
2
O and
denatured by adding 5.5 ml 3 M NaOH and incubating for
30 min at 42uC. Thereafter, 510 ml 2.3 M sodium bisulfite
(pH 5.0), 30 ml 10 mM hydroquinone and 65 ml ddH
2
O were
added, and the solution overlaid with mineral oil and held for 16h
at 55uC. The DNA was recovered using a WizardHDNA Clean-
Up System kit (Promega A7280), and a 90 ml aliquot treated with
10 ml 3 M NaOH for 15 min at 37uC, then neutralized by adding
70 ml 10 M ammonium acetate. Finally, the DNA was precipitat-
ed by adding 400 ml ethanol and 10 ml glycogen, and re-suspended
in 50 ml ddH
2
O to provide the template for a series of PCRs based
on the gene-specific primers listed in Table S2. A fragment of
Glyma20g32730 featuring many CG, CNG and CNN sites was
amplified from genomic DNA, then inserted into pMD18-T vector
and transferred into Dm- E. coli strain JM110. The plasmid was
released from the bacterial cells by the plasmid extraction kit
(TianGen. Cat. DP103-03) and treated with bisulfite in parallel
with the soybean genomic DNA as a control to monitor the
transformation efficiency of unmethylated cytosine to thymine.
The subsequent PCR consisted of 34–37 cycles of 94uC/30 s,
55uC/30 s, and 72uC/40 s. The resulting amplicons were purified
with a WizardHDNA Clean-Up System kit, ligated into the
pMD18-T vector (TaKaRa) and transferred into E. coli for
sequencing. Ten clones from each amplicon were sequenced.
Figure 2. Expression of 45 salinity inducible TFs in seedlings exposed to 5-ADC treatment. (a) GmAP2-DREBs, (b) GmMYBs, (c). GmNACs.
and (d) Gmb-ZIPs. M0-M72 refers to seedlings treated with water only for, respectively 0h, 12h, 24h, 48h and 72h, while A0-A72 refer to seedlings
exposed to 50 mM 5-ADC for 0h, 12h, 24h, 48h and 72h, respectively. Each gene-specific region was amplified by RT–PCR using the gene-specific
primers (Table S2). The TUBULIN gene (Genbank accession AY907703) was used as an internal control. The experiment was repeated three times with
similar result.
doi:10.1371/journal.pone.0041274.g002
Epigenetic Modifications Affect Salt Response
PLoS ONE | www.plosone.org 3 July 2012 | Volume 7 | Issue 7 | e41274
Figure 3. Methylation status of the promoter region of four salinity-responsive TFs in untreated (S0) and salinity-stressed (S1–S24)
seedlings (S1: 1h, S3: 3h, S6: 6h, S12: 12h, S24: 24h). (a) The black and white boxes indicate, respectively, exon and untranslated regions. The
short bars annotated with ‘‘I, II, III’’, ‘‘a’’ or b’’ indicate, respectively the sequences subjected to ChIP analysis, genomic bisulfite sequencing and those
used as probes for Southern blotting. The long vertical bars marked ‘‘c’’ display the distribution of CG dinucleotides (marked with red vertical lines),
and CNG (blue vertical lines) and CNN (black vertical lines) trinucleotides. The red vertical lines marked with a rectangular indicate CCGG sites
analyzed by Southern blotting. The thick black vertical lines represent the proportion of methylated cytosine. Ten positive clones from each gene’s
amplicon were sequenced. The data reflect the outcome of three independent experiments, and error bars represent standard error (SD). (b) The
efficiency of the bisulfite treatment to transform unmethylated cytosine to thymine. A fragment of Glyma20g32730 with numerous cytosines was
cloned into Dm- E. coli cells and the plasmid was treated with bisulfite in parallel with the soybean genomic DNA. All clones processed showed a
transformation rate .99.7%. (c) Methylation-sensitive DNA gel blot analysis of non-stressed (S0) and salinity-stressed seedlings (S1–S24). Genomic
DNA was digested to generate large fragments, then with one or other of the schizomers HpaII or MspI. Hybridization probes indicated. A DNA
fragment amplified from the probe sequence was used as a positive control (+), and ddH
2
O was used as a negative control (2).
doi:10.1371/journal.pone.0041274.g003
Epigenetic Modifications Affect Salt Response
PLoS ONE | www.plosone.org 4 July 2012 | Volume 7 | Issue 7 | e41274
The process was repeated three times using biologically indepen-
dent samples.
Chromatin immunoprecipitation (ChIP) assay
The ChIP protocol was modified from that of Johnson et al.
(2002). Briefly, 1 g of plant tissue was fixed by immersion in 1% v/
v formaldehyde under vacuum for 10 min. The extracted DNA/
protein complex was then sheared by sonication to a size range of
,100–1000 bp. After centrifugation, the complex was immuno-
precipitated by challenging with H3K9ac, H3K4me3 and
H3K9me2 antibodies (Millipore cat. 07–392, 07–473 and 05–
768R) at a titer of 1:100. The residual protein was degraded by the
addition of 10 ml (20 mg/ml) proteinase K, followed by a phenol/
chloroform extraction. A 2 ml aliquot of the final solution was used
as a template for qRT-PCR analysis as described above. A 1000x
diluted input DNA (Input) obtained from 500 ml of extract was
purified in parallel with the immunoprecipitated samples as a
control, and ChIP reactions were also performed in the absence of
antibody (No AB) to detect the occurrence of any non-specific
binding. Relative levels of H3K9 acetylation, H3K9 dimethylation
and H3K4 trimethylation were normalized to an internal control
(GenBank accession AY907703). The sequences of all PCR
primers used are given in Table S2. The mean and standard
deviation are shown for three independent ChIP experiments and
the significance of differences between means assessed with a ttest.
Methylation-sensitive Southern blot analysis
Genomic DNA (100 mg) extracted from both non-stressed and
salinity-stressed seedlings was treated for at least 6h with 100 U of
the appropriate restriction enzymes (EcoRV and NdeI for
Glyma11g02400;SacI for Glyma16g27950;BglII for Glyma20g30840;
SacI and BglII for Glyma08g41450) (TaKaRa) to generate large
fragments containing the target sequences. The digested DNA was
extracted by phenol/chloroform and divided into two equal
aliquots, one of which was treated with HpaII and the other with
its schizomer MspI [39]. The digested DNA was re-extracted,
electrophoretically separated through an 0.8% agarose gel and
transferred onto a Hybond N
+
membrane (Amersham). Probes for
each gene were designed to detect the methylation status within
the target sequence that analysed by genomic bisulfite sequencing.
About 3 mg of probe DNA was labeled using a DIG-High Prime
kit (Roche), and the subsequent hybridization and detection
procedure was performed using a DIG High Prime DNA Labeling
and Detection Starter Kit I (Roche), according to the manufac-
turer’s instructions. The positive control consisted of a 100x
diluted DNA fragment that amplified from the genomic DNA in
the same regions that prepared for probes, while the negative
control was ddH
2
O.
Results
The identification of salinity stress responsive TFs in
soybean
A set of differentially expressed genes were identified by
comparing the soybean Affymetrix microarray profiles generated
by probing with RNA extracted from salinity-stressed and non-
stressed plants. We mainly focused on the four groups of AP2/
EREB, bZIP, NAC and MYB transcription factors that have been
verified for salt stress in Arabidopsis or other plants. Of the 1,335
MYB, NAC, AP2/DREB and b-ZIP TFs represented on the
microarray, 49 appeared to be up-regulated (fold change of
hybridization signal .2, p,0.01) by salinity stress. These consisted
of 15 (of 448) GmMYBs, 9 (of 226) GmNACs, 16 (of 426) GmAP2/
DREBs and 9 (of 235) Gmb-ZIPs (Table S1). When the expression
of these 49 TFs was assayed by sqRT-PCR in both mock-stressed
(M0–M24) and salinity-stressed soybean seedlings (S1–S24) with
gene-specific primers (Table S2), 14 of the GmMYBs, 8 of the
GmNACs, 15 of the GmAP2/DREBs and 8 of the Gmb-ZIPs were
confirmed to be markedly induced by salinity stress (Figure 1).
Expression pattern analysis indicated that 19 of them were
strongly induced at a relatively early stage of exposure to salinity
(1–3h), while the others were induced somewhat later (6–24h).
Expression of the salinity induced TFs in the presence of
5-ADC
The expression of the 45 salinity-induced TFs was then
monitored in the mock treated (M0–M72) and the seedlings that
exposed to 5-ADC for various periods (A0–A72). As a result, ten of
the them showed higher levels of expression in treated (M0–M72)
than in mock-treated (A0–A72) seedlings; these ten TFs consisted
Figure 4. DNA methylation patterns in
Glyma11g02400
,
Glyma16g27950
,
Glyma08g41450
and
Glyma20g30840
in non-stressed (S0) and
salinity-stressed (S1 S24) seedlings. The left axis shows the percentage of methylated cytosines at each site (present as CG, CNG and CNN). The
data represent the mean of three biological replicates. Error bars represent standard errors.
doi:10.1371/journal.pone.0041274.g004
Epigenetic Modifications Affect Salt Response
PLoS ONE | www.plosone.org 5 July 2012 | Volume 7 | Issue 7 | e41274
of four GmMYBs(Glyma11g02400,Glyma07g30860,Glyma12g34650,
Glyma15g07230), one GmNAC (Glyma15g08480), four GmAP2/
DREBs(Glyma20g32730,Glyma20g30840,Glyma16g27950,Gly-
ma10g00980) and one Gmb-ZIP genes (Glyma8g41450) (Figure 2).
The expression level of nine of these TFs was very low for the first
12h of exposure, but thereafter rose substantially; the exception
was the Gmb-ZIP Glyma08g41450, the expression of which was
induced somewhat earlier.
Figure 5. Promoter methylation status in four salinity-responsive TFs in non-treated (A0) and 5-ADC treated seedlings (A12–A72).
For Figure legend please refer to Figure 3 legend.
doi:10.1371/journal.pone.0041274.g005
Epigenetic Modifications Affect Salt Response
PLoS ONE | www.plosone.org 6 July 2012 | Volume 7 | Issue 7 | e41274
Figure 6. Expression, DNA methylation and histone modification status of
Glyma11g02400, Glyma16g27950, Glyma08g41450
and
Glyma20g30840
in none treated (S0) and salinity-stressed (S1–S24) seedlings. (a) Relative H3K9 demethylation, acetylation and H3K4
Epigenetic Modifications Affect Salt Response
PLoS ONE | www.plosone.org 7 July 2012 | Volume 7 | Issue 7 | e41274
Methylation status as affected by salinity stress
To investigate the DNA methylation status of above candidate
genes under salinity stress, the sequence corresponding to the
translation start codon and the promoter region of the ten TFs was
subjected to bisulfite sequencing. First, the efficiency of the sodium
bisulfite treatment to convert cytosine to thymine was estimated.
The efficiency of the sodium bisulfite treatment to convert cytosine
to thymine in the cytosine rich segment of Glyma20g32730 was
estimated to be 99.7%. Bisulfite sequencing result indicated that
the Glyma11g02400, Glyma08g41450, Glyma16g27950 and Gly-
ma20g30840 promoters all appeared to be differentially methylated
by the imposition of salinity stress (Figure 3a, b, Figure S1), but
those of the other six genes were largely non-methylated
(Figure S2). In the Glyma11g02400 promoter from position 2518
to 2274, most of the cytosines were demethylated following
exposure to salinity stress for 1–24h (Figure 3a). In the immediate
downstream region of the Glyma16g27950 transcription start codon
(+24 to +233), about 35% of the cytosines were methylated both
before the salinity stress was imposed and for the first three hours
of stress, but thereafter only few methylated cytosines remained
(Figure 3a). In the Glyma20g30840 promoter region 1 (287 to
+163), 51% of the cytosines were methylated prior to exposure to
salinity stress, but this proportion fell to 27% after 1h, 12% after
3h, and to even lower levels as the stress was prolonged further
(Figure 3a); meanwhile in region 2 of the same promoter (2163 to
2405), 42% of the cytosines were methylated at the start of the
stress period, and this proportion hardly altered thereafter
(Figure 3a). In the Glyma08g41450 region immediately downstream
of the transcription start codon (+24 to +233), 35% of the cytosines
were methylated prior to the imposition of stress, and the same
proportion was maintained throughout (Figure 3a).
DNA methylation-sensitive Southern blotting was applied to
verify these observations. The restriction fragments including
Glyma11g02400,Glyma16g27950 and Glyma20g30840 were more
readily digested by HpaII after 6h of salinity stress than the same
samples obtained from non-stressed seedlings, which consistent
with a reduction in global cytosine methylation caused by salinity
stress analyzed by bisulfite sequencing (Figure 3a, c). The sequence
surrounding Glyma08g41450 was not digestible by HpaII, but
several small restriction fragments were generated by MspI
digestion, suggesting that CCGG sites in the region of this TF
were hypermethylated in both non-stressed and stressed seedlings
(Figure 3c). Thus the Southern blotting outcomes were in general
consistent with the bisulfite sequence data.
The analysis of DNA methylation pattern of them indicated that
methylation was affecting either CG dinucleotides or CNG/CNN
trinucleotides under salinity stressed process (Figure 4). In the
Glyma11g02400 promoter, 98% of the CG’s, 60% of the CNG’s
and 6% of the CNN’s were methylated in the non-stressed
seedlings, but almost all the CG’s, CNG’s and CNN’s were
demethylated in plants exposed to salinity stress for more than 6h
(Figure 4). For Glyma16g27950, some 80% of the CG’s, 72% of the
CNG’s and 4% of the CNN’s were methylated in non-stressed
seedlings and those during the early phase (1–3h) of the salinity
treatment, but by 6h a significant fall in CG and CNG methylation
was observed (Figure 4). Within region 1 of the Glyma20g30840
promoter, 95% of the CG’s, along with 40% of the CNG’s and 4%
of the CNN’s, were methylated at 3h after the imposition of stress,
but by 6h, a marked reduction in CG and CNG methylation had
occurred. CG, CNG and CNN methylation in the Glyma08g41450
promoter was unaffected by salinity stress (Figure 4). Clearly, DNA
methylation in Glyma11g02400, Glyma16g27950 and Gly-
ma20g30840 (region 1) varied over the period of the salinity stress
episode.
Methylation status of Glyma11g02400, Glyma08g41450,
Glyma16g27950 and Glyma20g30840 as affected by the
presence of 5-ADC
To identify whether the up-regulation of these four genes were
related with cytosine demethylation under 5-ADC treatment, the
effect on the DNA methylation status of the four responsive TFs in
plants treated with 5-ADC was analyzed using genomic bisulfite
sequencing. The relevant test for the transformation efficiency of
unmethylated cytosine to thymine using Glyma20g32730 is
illustrated. As a result, all four TFs were hypermethylated in
non-treated seedlings; after a 24h exposure to 5-ADC, some
evidence of demethylation was obtained, but from 48h onwards it
was clear that a substantial level of demethylation had occurred
(Figure 5a, b). This observation was supported by a DNA
methylation-sensitive DNA gel blot (Figure 5c). All of the four
TFs showed an increased digestion with HpaII after salinity stress
for more than 48h, suggesting a reduction of cytosine methylation
in 5-ADC stressed seedlings (Figure 5c).
Histone modification of hypermethylated genes induced
by salinity stress
The histone content (H3K4me3, H3K9ace and the inactive
H3K9me2) of the four TFs (Glyma11g02400,Glyma08g41450,
Glyma16g27950 and Glyma20g30840) which responded to salinity
stress by altering their methylation status was then examined,
using a combination of ChIP and qRT-PCR (Figure 6a). An
unmethylated gene Glyma20g32730 was also analysed in parralle
with them as a control (Figure S3). When Glyma11g02400 was
induced by salinity stress, a significant increase in H3K4me3
(regions I, II and III) and a decrease in H3K9me2 (regions I and
II) was observed, but the H3K9ace sites remained unmodified
(Figure 6a). Within Glyma20g30840 (regions II and III) and
Glyma08g41450 (region III), a high level of H3K9me2 and a low
level of H3K4me3 and H3K9ac was present in both non-stressed
seedlings and those sampled during the early phase (1–3h) of
salinity stress; at later time points (6–24h), a significant decrease in
H3K9me2 and increase of H3K4me3 and H3K9ac content was
observed (Figure 6a). A similar H3K4me3, H3K9ac or H3K9me2
signal was detected in all three regions of the Glyma16g27950
promoter (Figure 6a). Thus, like the DNA methylation, histone
modification was also subject to dynamic change during the course
of the salinity stress episode.
Changes of epigenetic modification in regulating the TFs
expression during salinity stress
The expression of Glyma11g02400 was low in non-stressed
seedlings, while its promoter was hypermethylated and was highly
enriched for H3K9me2 and depleted for H3K4me3 (Figure 6a, b,
trimethylation content (ChIP assay). A 1:1,000 dilution of input DNA (Input) served as a control for PCR amplifications and the ChIP reactions carried
out in the absence of antibody (N0 AB). Relative H3K9 acetylation, H3K9 dimethylation and H3K4 trimethylation were determined by qRT-PCR and
normalized to an internal control TUBULIN gene (Genbank accession AY907703). Data represent the mean of three biological replicates. Asterisks
indicate means differing significantly from the S0 situation. Error bars represent standard errors. *P,0.05, **P,0.01. (b) Gene expression (qRT-PCR)
profiles. (c) Cytosine methylation level (bisulfite sequencing).
doi:10.1371/journal.pone.0041274.g006
Epigenetic Modifications Affect Salt Response
PLoS ONE | www.plosone.org 8 July 2012 | Volume 7 | Issue 7 | e41274
c). When the seedlings were exposed to salinity stress, its
expression rose rapidly, while its promoter became gradually
demethylated, the level of H3K9me2 fell and that of H3K4me3
increased (Figure 6a, b, c). Glyma20g30840 behaved similarly, but
the establishment of H3K4me3, DNA demethylation and gene
expression were not contemporaneous. Demethylation was
noticeable within 1h of the imposition of salinity stress, but the
establishment of H3K4me3 did not occur until 3h and the up-
regulation of expression only at 6h (Figure 6a, b, c). Similarly,
Glyma08g41450 was up-regulated by 12h after the start of the stress
episode, but its promoter was hypermethylated throughout.
Between 6h and 24h after the stress had begun, the level of
H3K9me2 fell and that of H3K4me3 and H3K9ac rose (Figure 6a,
b, c). Glyma16g27950 expression was repressed in non-stressed
seedlings and during the early phase of the stress episode, when its
promoter was hypermethylated. Later the TF was gradually
induced and its promoter demethylated, while the level of
enrichment of H3K9me2, H3K4me3 and H3K9ac did not
change. A correlation analysis of their expression, methylation
levels and histone modifications indicated that up-regulation of
Glyma11g02400 was associated with a decreased level of DNA
methylation (r = 20.89), H3K9me2 (r = 20.965, on average of
regions II and III) and an increased level of H3K4me3 (r =0.7, on
average of regions I and II) (Figure 6; Table S3). The expression of
Glyma20g30840 correlated negatively with DNA methylaion
(r = 20.78), H3K9me2 (r = 20.93 on average of regions II and
III) and positively with H3K9ac (r = 0.96) and H3K4me3 (r = 0.93
on average) (Figure 6; Table S3). Up-regulation of Glyma16g27950
just correlated negatively with DNA methylation (r = 20.97), did
not with histone modifications, while the up-regulation of
Glyma08g41450 correlated negatively with H3K9me2 (r = 20.67)
and positively with H3K9ac (r = 0.84) and H3K4me3 (r = 0.78),
did not with DNA methylation during salinity stress (Figure 6;
Table S3). Therefore, the histone accumulation/depletion and/or
cytosine methylation appears to underlie the activation by salinity
stress of these four TFs.
Discussion
TF transcription can be influenced by both DNA
methylation and/or histone modification in a region
specific manner
The regulation of genes via cytosine methylation and histone
modification is a well recognized component of the plant stress
response [14,34,40,41]. Both these epigenetic modifications are
region-specific, and can be dynamic over time [42,43,44,45]. Here
we have identified a set of ten salinity-induced, cytosine
methylation-dependent TFs, three of which displayed the expected
relationship between promoter cytosine methylation and gene
expression during the salinity stress process, but one of which
(Glyma08g41450) remained up-regulated even though it was in a
highly methylated state (Figure 1, Figure 3a). A similar unexpected
relationship has been noted for an embryogenesis-related gene in
carrot [42]. Many genes are expressed despite their promoter
region being highly methylated, so it seems probable that for
Glyma08g41450, its up-regulation in plants exposed to salinity stress
is independent of DNA methylation. One region of the
Glyma20g30840 promoter was hypermethylated in both stressed
and non-stressed plants, while its neighbouring region responded
to the stress by a reduction in methylation (Figure 3a). This
behaviour provides an example of region-specific regulation of
methylation. The gradual up-regulation of Glyma11g02400 in
salinity-stressed seedlings was accompanied by a decrease in CG,
CNG and CNN methylation (Figure 4), while the rapid up-
regulation of Glyma20g30840 and Glyma16g27950 was accompa-
nied by a decrease in only CG and CNG methylation (Figure 4),
suggesting heterogeneity in the genome for gene expression
regulation via DNA methylation.
Histone modification provides a second major mechanism of
epigenetic control over gene expression [46]. In a range of A.
thaliana stress-responsive genes, salinity stress has been shown to
increase trimethylation at H3K4 and decrease H3K9 demethyl-
ation [28]. The up-regulation of Glyma20g30840 and Gly-
ma08g41450 during the course of the salinity stress episode may
have been achieved by the depletion of H3K9me2 and the
enrichment of H3K4me3 and H3K9ac in regions II and III of
Glyma20g30840 and in region III of Glyma08g41450 (Figure 6a, b).
The up-regulation of Glyma11g02400 may have been brought
about by the depletion of H3K9me2 and the enrichment of
H3K4me3 in regions I, II and III (Figure 6a, b). Thus, as for DNA
methylation, the effect of salinity stress on histone modification
appears to be heterogeneous across the genome.
Transcriptional activation, DNA methylation and histone
modification are not simultaneous events
There was a distinct time lag between transcriptional activation,
DNA methylation and/or histone modification of Glyma11g02400,
Glyma20g30840 and Glyma08g41450. Thus, most of the cytosine
content of the Glyma11g02400 promoter was demethylated very
soon (within 1h) of the imposition of stress, but there was no
evidence for the TF’s activation before 3h (Figure 6b, c). Similarly,
Glyma20g30840 was progressively demethylated over the full 24h
of the stress episode, but its up-regulation was complete within 6h
(Figure 6b, c). In Glyma11g02400,Glyma20g30840 and
Glym08g41450, the H3K4me3 content was already increasing by
1h, but the up-regulation of TF expression occurred substantially
later (Figure 6a, b). Similar time lags have been noted for the
build-up of H3K4me3 within the coding regions of the A. thaliana
drought-related genes RD29A and RAP2.4 during drought stress
[36]. Furthern more, the DNA demethylation of Glyma20g30840
was prior to the erasure or establishment of H3K9me2, H3K4me3
and H3K9ac (Figure 6a, c) and all the three genes (Gly-
ma11g02400, Glyma20g30840 and Glym 08g41450) show an ealier
establishment of H3K4me3 than H3K9me2 and H3K9ac
(Figure 6a). Suggesting that there were also a time lage between
the DNA demethylation and establishment of histone modifaci-
tions.
The interplay between DNA methylation and histone
modification in the context of patterns of gene
expression
In A. thaliana, it has been demonstrated that the loss of CG
methylation in met1 plants has a large effect on H3K9me2 content,
leading to the idea that cytosine methylation can influence H3K9
modification [17,47,48,49]. H3K9me2, mediated by KYP/
SUVH4 and SUVH2, is also known to direct non-CG methylation
[39,50]. In the ‘two-step’ hypothesis for the regulation of
transcription, CG methylation directs H3K9 methylation and
H3K9 methylation recruits non-CG methylation [51]. Moreover,
hypermethylation of the stress inducible genes in Arabidopsis
correlated with the enrichment of H3K9me2 and depletion of
H3K9ac histones under salt stress conditions [16]. In this study,
the behaviour of Glyma11g02400 and Glyma20g30840 was consis-
tent with this model; as CG was progressively demethylated in
these TFs, the content of H3K9me2, H3K4me3 and/or H3K9ac
rose and meanwhile a lower level of non-CG methylation was
observed (Figure 6a, Figure 4). Note, however, that for
Epigenetic Modifications Affect Salt Response
PLoS ONE | www.plosone.org 9 July 2012 | Volume 7 | Issue 7 | e41274
Glyma16g27950, while salinity stress led to a noticeable demeth-
ylation of the DNA, it had little effect on the level of histone
modification (Figure 6a, c). The Glyma08g41450 promoter was
hypermethylated throughout the salinity stress period, while
H3K9me2 was depleted and H3K4me3 and H3K9ac accumulat-
ed (Figure 6a, c). For these two TFs, therefore, the evidence is that
DNA methylation had no influence on histone modification,
consistent with the behaviour of the A. thaliana genes TOUSLED
and RPA2 [52,53].
Supporting Information
Figure S1 Monitoring the non-methylated cytosine to
thymine transformation efficiency of the bisulfite se-
quencing procedure. A fragment of Glyma20g32730 with
numerous cytosines was cloned into Dm- E. coli cells and the
plasmid was treated with bisulfite. Nine positive clones of this gene
from the bisulfite treated DNA were sequenced and compared
with the untreated DNA sequence.
(PDF)
Figure S2 Genomic bisulphite sequencing of six un-
methylated genes in plants not exposed to salinity (S0)
and those stressed for 1–24 h (S1–S24). Ten clones of each
gene were sequenced. The short thick bars indicate the regions
subjected to sequencing, and the long vertical ones show the
distribution of CG dinucleotides (black circles), and CNG
(triangles) and CNN (unmarked) trinucleotides. The black vertical
line represents the proportion of methylated sites. Data represent
the mean of three biological replicates.
(PDF)
Figure S3 ChIP analysis to assess the unmethylated
gene Glyma20g32730’s H3K9me2, H3K9ac and
H3K4me3 content in plants challenged with salinity.
The short bars marked ‘‘a’’ indicate regions subjected to genomic
bisulfite sequencing (+1to+393); ‘‘I, II and III’’ indicate the
regions subjected to ChIP analysis.
(PDF)
Table S1 The set of TFs identified by microarray
analysis to be inducible by salinity stress.
(DOC)
Table S2 Sequences of primers employed in this
research.
(DOC)
Table S3 Correlation analysis among methylation lev-
els, gene expression and histone modifications of the
four TFs during salinity stress. NC: no correlation;
*P
,
0.05, **P
,
0.01.
(DOC)
Acknowledgments
We thank Dr. Robert Koebner (UK) for English editing of the manuscript.
Author Contributions
Conceived and designed the experiments: FX. Performed the experiments:
YS DJ SL PW QL. Analyzed the data: YS. Contributed reagents/
materials/analysis tools: FX. Wrote the paper: YS.
References
1. Phang TH, Shao G, Lam HM (2008) Salt tolerance in Soybean. J Integr Plant
Biol 50: 1196–1212.
2. Cheong YH, Moon BC, Kim JK, Kim CY, Kim MC, et al. (2003) BWMK1, a
rice itogen-activated protein kinase, locates in the nucleus and mediates
pathogenesis-related gene expression by activation of a transcription factor.
Plant Physiol 132: 1961–1972.
3. Choi HI, Hong JH, Ha JO, Kang JY, Kim SY (2000) ABFs, a family of ABA-
responsive element binding factors. J Biol Chem 275: 1723–1730.
4. Uno Y, Furihata T, Abe H, Yoshida R, Shinozaki K, et al. (2000) Arabidopsis
basic leucine zipper transcription factors involved in an abscisic acid-dependent
signal transduction pathway under drought and high-salinity conditions. Proc
Natl Acad Sci U S A 97: 11632–11637.
5. Sakuma Y, Liu Q, Dubouzet JG, Abe H, Shinozaki K, et al. (2002) DNA-
binding specificity of the ERF/AP2 domain of Arabidopsis DREBs, transcrip-
tion factors involved in dehydration- and cold inducible gene expression.
Biochem Biophys Res Commun 290: 998–1009.
6. Abe H, Urao T, Ito T, Seki M, Shinozaki K, Yamaguchi-Shinozaki K (2003)
Arabidopsis AtMYC2 (b-HLH) and AtMYB2 (MYB) function as transcriptional
activators in abscisic acid signalling. The Plant Cell 15: 63–78.
7. Chen M, Xu Z, Xia L, Li L, Cheng X, et al. (2009) Cold-induced modulation
and functional analyses of the DRE-binding transcription factor gene,
GmDREB3, in soybean (Glycine max L.). J Exp Bot 60: 121–35.
8. Bu QY, Jiang HL, Li CB, Zhai QZ, Zhang J, et al. (2008) Role of the
Arabidopsis thaliana NAC transcription factors ANAC019 and ANAC055 in
regulating jasmonic acid-signaled defense responses. Cell Res 18: 756–767.
9. Liao Y, Zou HF, Wei W, Hao YJ, Tian AG, et al. (2008a) Soybean GmbZIP44,
GmbZIP62 and GmbZIP78 genes function as negative regulator of ABA
signaling and confer salt and freezing tolerance in transgenic Arabidopsis. Planta
228: 225–240.
10. Liao Y, Zou HF, Wang HW, Zhang WK, Ma B, Z, et al. (2008b) Soybean
GmMYB76, GmMYB92, and GmMYB177 genes confer stress tolerance in
transgenic Arabidopsis plants. Cell Research 18: 1047–1060.
11. Chen M, Wang QY, Cheng XG., Xu ZS, Li LC, et al. (2007) GmDREB2, a
soybean DRE-binding transcription factor, conferred drought and high-salt
tolerance in transgenic plants. Biochem Biophys Res Commun 353: 299–305.
12. Hao YJ, Wei W, Song QX, Chen HW, Zhang YQ, et al. (2011) Soybean NAC
transcription factors promote abiotic stress tolerance and lateral root formation
in transgenic plants. Plant J 68: 302–313.
13. Zhu JK (2008) Epigenome sequencing comes of age. Cell 133: 395–397.
14. Chinnusamy, Gong Z, Zhu JK (2008) ABA-mediated epigenetic processes in
plant development and stress responses. J Integr Plant Biol 50: 1187–1195.
15. Boyko A, Kovalchuk I (2008) Epigenetic control of plant stress response. Environ
Mol Mutagen 49: 61–72.
16. Bilichak A, Ilnystkyy Y, Hollunder J, Kovalchuk I (2012) The progeny of
Arabidopsis thaliana plants exposed to salt exhibit changes in DNA methylation,
histone modifications and gene expression. PLoS One 7(1): e30515.
17. Tariq M, Paszkowski J (2004) DNA and histone methylation in plants. Trends
Genet 20: 244–252.
18. Peters AH, Schubeler D (2005) Methylation of histones: playing memory with
DNA. Curr Opin Cell Biol 17: 230–238.
19. Fuchs J, Demidov D, Houben A, Schubert I (2006) Chromosomal histone
modification patterns – from conservation to diversity. Trends Plant Sci 11: 199–
208.
20. Paszkowsk i J, Whitham SA (2001) Gene silencing and DNA methylation
processes. Curr Opin Plant Biol 4: 123–129.
21. Bird A (2002) DNA methylation patterns and epigenetic memory. Genes &
Development 16: 6–21.
22. Finnegan EJ, Genger RK, Peacock WJ, Dennis ES (1998) DNA methylation in
plants. Annu Rev Plant Physiol Plant Mol Biol 49: 223–247.
23. Zhang XY, Yazaki J, Sundaresan A, Cokus S, Chan SW, et al. (2006) Genome-
wide high-resolution mapping and functional analysis of DNA methylation in
Arabidopsis. Cell 126: 1189–1201.
24. Vaillant I, Paszkowski J (2007) Role of histone and DNA methylation in gene
regulation. Curr Opin Plant Biol 10: 528–533.
25. Lachner M, O’Sullivan RJ, Jenuwein T (2003) An epigenetic road map for
histone lysine methylation. J Cell Sci 116: 2117–2124.
26. Earley K, Lawrence RJ, Pontes O, Reuther R, Enciso AJ, et al. (2006) Erasure of
histone acetylation by Arabidopsis HDA6 mediates large-scale gene silencing in
nucleolar dominance. Genes Dev 20: 1283–1293.
27. Chen ZJ, Tian L (2007) Roles of dynamic and reversible histone acetylation in
plant development and polyploidy. Biochim Biophys Acta 1769: 295–307.
28. Chen LT, Luo M, Wang YY, Wu K (2010) Involvement of Arabidopsis histone
deacetylase HDA6 in ABA and salt stress response. J Exp Bot 61: 3345–3353.
29. Tamaru H, Selker EU (2001) A histone H3 methyltransferase controls DNA
methylation in Neurospora crassa. Nature 414: 277–283.
30. Jackson JP, Johnson L, Jasencakova Z, Zhang X, PerezBurgos L, et al. (2004)
Dimethylation of histone H3 lysine 9 is a critical mark for DNA methylation and
gene silencing in Arabidopsis thaliana. Chromosoma 112: 308–315.
31. Johnson L, Cao X, Jacobsen S (2002) Interplay between Two Epigenetic Marks:
DNA Methylation and Histone H3 Lysine 9 Methylation. Curr Biol 12: 1360–
1367.
Epigenetic Modifications Affect Salt Response
PLoS ONE | www.plosone.org 10 July 2012 | Volume 7 | Issue 7 | e41274
32. Fahrner JA, Eguchi S, Herman JG, Baylin SB (2002) Dependence of histone
modifications and gene expression on DNA hypermethylation in cancer. Cancer
Res 62: 7213–7218.
33. Steward N, Ito M, Yamaguchi Y, Koizumi N, Sano H (2002) Periodic DNA
methylation in maize nucleosomes and demethylation by environmental stress.
J Biol Chem 277: 37741–37746.
34. Choi CS, Sano H (2007) Abiotic-stress induces demethylation and transcrip-
tional activation of a gene encoding a glycerophosphodiesterase-like protein in
tobacco plants. Mol Genet Genomics 277: 589–600.
35. Tsuji H, Saika H, Tsutsumi N, Hirai A, Nakazono M (2006) Dynamic and
reversible changes in histone H3-Lys4 methylation and H3 acetylation occurring
at submergence-inducible genes in rice. Plant Cell Physiol 47: 995–1003.
36. Kim JM, To TK, Ishida J, Morosawa T, Kawashima M, et al. (2008) Alterations
of lysine modifications on histone H3 N-tail under drought stress conditions in
Arabidopsis thaliana. Plant Cell Physiol 49: 1580–1588.
37. Smyth GK, Yang YH, Speed TP (2003) Statistical issues in microarray data
analysis. In: Brownstein MJ, Khodu rsky AB, eds. Functional Genomics:
Methods and Protocols. Volume 224, 111–136 Humana Press, Totowa.
38. Livak KJ, Schmittgen TD (2001) Analysis of relative gene expression data using
real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods 25:
402–408.
39. Jackson JP, Lindroth AM, Cao X, Jacobsen SE (2002) Control of CpNpG DNA
methylation by the KRYPTONITE histone H3 methyltransferase. Nature 416:
556–560.
40. Zhu J, Jeong JC, Zhu Y, Sokolchik I, Miyazaki S, et al. (2007) Involvement of
Arabidopsis HOS15 in histone deacetylation and cold tolerance. Proc Natl Acad
Sci U S A 105, 4945–4950.
41. Agius F, Kapoor A, Zhu JK (2006) Role of the Arabidopsis DNA glycosylase/
lyase ROS1 in active DNA demethylation. Proc Natl Acad Sci U S A 103:
11796–11801.
42. Shibukawa T, Yazawa K, Kikuchi A, Kamada H (2009) Possible involvement of
DNA methylation on expression regulation of carrot LEC1 gene in its 59-
upstream region. Gene 437: 22–31.
43. Pokholok DK, Harbison CT, Levine S, Cole M, Hannett NM, et al. (2005)
Genome-wide map of nucleosome acetylation and methylation in yeast. Cell
122: 517–527.
44. Millar CB, Grunstein M (2006) Genome-wide patterns of histone modifications
in yeast. Nat Rev Mol Cell Biol 7: 657–666.
45. Li X, Wang X, He K, Ma Y, Su N, et al. (2008) High-resolution mapping of
epigenetic modifications of the rice genome uncovers interplay between DNA
methylation, histone methylation, and gene expression. Plant Cell 20: 259–276.
46. Egger G., Liang GN, Aparicio A, Jones PA (2004) Epigenetics in human disease
and prospects for epigenetic therapy. Nature 429: 457–463.
47. Mathieu O, Probst AV, Paszkowski J (2005) Distinct regulation of histone H3
methylation at lysines 27 and 9 by CpG methylation in Arabidopsis. EMBO J
24: 2783–2791.
48. Tariq M, Saze H, Probst AV, Lichota J, Habu Y, et al. (2003) Erasure of CpG
methylation in Arabidopsis alters patterns of histone H3 methylation in
heterochromatin. Proc Natl Acad Sci U S A 100: 8823–8827.
49. Soppe WJ, Jasencakova Z, Houben A, Kakutani T, Meister A, et al. (2002) DNA
methylation controls histone H3 lysine 9 methylation and heterochromatin
assembly in Arabidopsis. EMBO J 21: 6549–6559.
50. Malagnac F, Bartee L, Bender J (2002) An Arabidopsis SET domain protein
required for maintenance but not establishment of DNA methylation. EMBO J
21: 6842–6852.
51. Naumann K, Fischer A, Hofmann I, Krauss V, Phalke S, et al. (2005) Pivotal
role of AtSUVH2 in heterochromatic histone methylation and gene silencing in
Arabidopsis. EMBO J 24: 1418–1429.
52. Xia R, Wang J, Liu C, Wang Y, Zhai J, et al. (2006) ROR1/RPA2A, a putative
replication protein A2, functions in epigenetic gene silencing and in regulation of
meristem development in Arabidopsis. Plant Cell 18: 85–103.
53. Kapoor A, Agarwal M, Andreucci A, Zheng X, Gong Z, et al. (2005) Mutations
in a conserved replication protein suppress transcriptional gene silencing in a
DNA-methylation-independent manner in Arabidopsis. Curr Biol 15: 1912–
1918.
Epigenetic Modifications Affect Salt Response
PLoS ONE | www.plosone.org 11 July 2012 | Volume 7 | Issue 7 | e41274
... On the contrary, four TFs showed induced response under salinity stress in soybean, and of these three showed demethylations in CG and non-CG contexts and also active enrichment of histone marks (H3K4me3 and H3K9ac) along with a reduction in the repressive mark H3K9me2. Thus, the possible interplay was recorded between methylation of DNA and histone modifications when exposed to salinity stress (Song et al. 2012). In the plants of salinity-stressed soybean, cross talk has been reported between histone methylation and acetylation (Liu et al. 2010;Stępiński 2012). ...
... Soybean Demethylation Song et al. (2012) 19. ...
... Soybean Hypomethylation Song et al. (2012) 20. ...
Chapter
Epigenetic modifications are known to alter the activation pattern of some genes and not the per se DNA sequence. Stress to the plant causes epigenetic alterations in the plant either as hyper- or as hypo-methylation of certain DNA sequences. To overcome or to counter the various abiotic stress conditions, the plant’s defense machinery including cellular signaling pathways gets regulated by several stress-responsive genes, which in turn are regulated by various mechanisms including DNA and chromatin modifications, and also through different small RNA-based mechanisms. There is a sudden spurt in the epigenetic studies aiming to find their role in the imposition of various types of abiotic stress tolerance in different plant species, mainly due to the quick advancements in the high-throughput NGS technologies. Many reports associating the DNA methylation response with that of various abiotic stress adaptations are available in many legume species like soybean, chickpea, pigeon pea, Medicago, lotus, peanut, and common beans using these techniques. These legumes have shown tolerance to several abiotic stresses because of unique epigenetic variations, which are present in the natural populations. Understanding the epigenetic mechanism regulating the tolerance to the abiotic stresses will help plant breeders in the development of more resilient and climate-smart varieties, giving higher yields under varied abiotic stresses. This chapter covers the current status of a novel and promising field of epigenetics in legume crops, especially for the imposition of different abiotic stress tolerance.
... Importantly, findings indicate that under certain stressful environments, active expression of stressrelated genes necessitates eliminating repressive methylation marks and the addition of active methylation marks (Sun et al., 2019). For instance, salinity stress results in the reduction of suppressive histone marks including H3K27me and 3H3K9me2, coupled with the accrual of functional methylation marks such as H3K4me3 on salt-tolerance genes (Song et al., 2012;Paul et al., 2017). In soybean, the deactivation of genes in salt-exposed plants firmly correlates with the H3K27me3 establishment under de novo conditions in the coding regions or promoters not present in control plants (Sun et al., 2019). ...
Article
The severe impact of drought and salinity on plant productivity presents a significant threat to worldwide food security. Plants exhibit the capacity to sense stimuli in their environment and adjust defense mechanisms through diverse regulatory networks to cope with abiotic stress. The complexities of drought and salinity tolerances can be deconstructed into contributing factors and mechanisms, classified under two categories: genetics and epigenetics. Epigenetic mechanisms play a role in partially attributing crop adaptation to the most formidable drought and salinity stresses. This review highlights the latest and noteworthy findings regarding crop epigenetic responses to abiotic stress signals, particularly those pertinent to drought and salinity tolerance.
... TFs are key proteins that regulate complex mechanisms in plants in response to various stresses. They help plants adapt and survive adverse environmental conditions by regulating gene expression [46]. The TFs identified in our study were distributed in 58 different families, including those known to be modulated by salt stress such as C2H2 [47], ERF [48], MYB [49], WRKY [50], and bHLH [51] (Table S13). ...
Article
Full-text available
Salt stress is a significant abiotic stress that reduces crop yield and quality globally. In this study, we utilized RNA sequencing (RNA-Seq) to identify differentially expressed genes (DEGs) in response to salt stress induced by gamma-ray irradiation in a salt-tolerant soybean mutant. The total RNA library samples were obtained from the salt-sensitive soybean cultivar Kwangan and the salt-tolerant mutant KA-1285. Samples were taken at three time points (0, 24, and 72 h) from two tissues (leaves and roots) under 200 mM NaCl. A total of 967,719,358 clean reads were generated using the Illumina NovaSeq 6000 platform, and 94.48% of these reads were mapped to 56,044 gene models of the soybean reference genome (Glycine_max_Wm82.a2.v1). The DEGs with expression values were compared at each time point within each tissue between the two soybeans. As a result, 296 DEGs were identified in the leaves, while 170 DEGs were identified in the roots. In the case of the leaves, eight DEGs were related to the phenylpropanoid biosynthesis pathway; however, in the roots, Glyma.03G171700 within GmSalt3, a major QTL associated with salt tolerance in soybean plants, was differentially expressed. Overall, these differences may explain the mechanisms through which mutants exhibit enhanced tolerance to salt stress, and they may provide a basic understanding of salt tolerance in soybean plants.
... As the study of epigenetics has advanced, it has provided novel directions to drive plant-breeding strategies by exploiting epigenetic variation and/or manipulating the epigenome to improve adaptation to various environmental stresses and ensure yield and quality (Rodrıǵuez Loṕez and Wilkinson, 2015;Tirnaz and Batley, 2019;Pecinka et al., 2020;Gupta and Salgotra, 2022) (Figure 2). Indeed, studies have revealed the relevance of epigenetic regulation of stress response in many model and crop species such as A. thaliana (e.g., Tricker et al., 2012), barley (e.g., Konate et al., 2018), maize (e.g., Steward et al., 2002), rice (e.g., Zheng et al., 2017), soybean (e.g., Song et al., 2012), tomato (e.g., Gonzaĺez et al., 2013), and wheat (e.g., Wang et al., 2016). Some additional examples include: Lämke et al. (2016) have described the methylation of histone H3 lysine 4 (H3K4) is involved in the heat stress-induced genes. ...
Article
Full-text available
Climate change is expected to increase the occurrence of extreme environmental conditions. Viticulture, as agriculture in general, is highly dependent on climatic conditions, not only for yield but also for fruit quality, which is the most important factor affecting produce value at the farm-gate. This demands the development of novel plant breeding techniques that will lead to the accelerated production of more resilient grape varieties, as conventional breeding programs for perennials are often prolonged. Recent research has suggested that environmental conditions can activate a memory of stress that could result in a primed response to subsequent stress events. This is a process capable of increasing plant’s resilience to abiotic stimuli, allowing plants to better adapt to extreme environmental conditions. While the effect of priming has been observed in many plants, the underlying mechanisms are puzzling and seldom studied in perennial crops. A large body of research has been developed in the last decade linking response to stress, stress priming, and memory of stress with epigenetic mechanisms. This understanding of plant epigenetics has opened the door to the application of epigenetics to crop improvement, such as the use of epigenetic breeding for the generation of more resilient crops. Perennial crop agriculture in general, and viticulture in particular, would benefit from more in-depth knowledge on epigenetic memory of stress.
... Up-regulation of cell-wall-related genes such as ZmXYLOGLUCAN ENDOTRANSGLUCOSYLASE/HYDROLASE1 and ZmEXPANSIN B2 are linked with acetylation of H3K9, which is a histone protein (H3) with lysine (K) on the 9th position, in both the promoter and coding region. Mutant gcn5 exhibits salt sensitive response in maize because of the cell wall integrity [207]. In another study, Arabidopsis GCN5 plays an important role in heat tolerance through H3K9/k14 acetylation process in the promoter sequence of the ULTRAVIOLET HYPERSENSITIVE6 and, HEAT SHOCK TRANSCRIPTION FACTOR A3 genes [208]. ...
Article
Full-text available
The present day’s ongoing global warming and climate change adversely affect plants through imposing environmental (abiotic) stresses and disease pressure. The major abiotic factors such as drought, heat, cold, salinity, etc., hamper a plant’s innate growth and development, resulting in reduced yield and quality, with the possibility of undesired traits. In the 21st century, the advent of high-throughput sequencing tools, state-of-the-art biotechnological techniques and bioinformatic analyzing pipelines led to the easy characterization of plant traits for abiotic stress response and tolerance mechanisms by applying the ‘omics’ toolbox. Panomics pipeline including genomics, transcriptomics, proteomics, metabolomics, epigenomics, proteogenomics, interactomics, ionomics, phenomics, etc., have become very handy nowadays. This is important to produce climate-smart future crops with a proper understanding of the molecular mechanisms of abiotic stress responses by the plant’s genes, transcripts, proteins, epigenome, cellular metabolic circuits and resultant phenotype. Instead of mono-omics, two or more (hence ‘multi-omics’) integrated-omics approaches can decipher the plant’s abiotic stress tolerance response very well. Multi-omics characterized plants can be used as potent genetic resources to incorporate into the future breeding program. For the practical utility of crop improvement, multi-omics approaches for particular abiotic stress tolerance can be combined with genome-assisted breeding (GAB) by being pyramided with improved crop yield, food quality and associated agronomic traits and can open a new era of omics-assisted breeding. Thus, multi-omics pipelines together are able to decipher molecular processes, biomarkers, targets for genetic engineering, regulatory networks and precision agriculture solutions for a crop’s variable abiotic stress tolerance to ensure food security under changing environmental circumstances.
... There are many reports on MYB TFs' regulation of abiotic stress in the plant [17,30,69,70,73], but its link with epigenetic control is rarely reported [30,31,74]. As CG or CpG is the most abundant source of DNA methylation, methylation-sensitive amplification polymorphism and methylation-specific PCR (MSP) are two methods that have been applied to methylated genes [75,76]. ...
Article
Full-text available
Background: Chrysanthemum is a popular ornamental plant worldwide. MYB (v-myb avian myeloblastosis viral oncogene homolog) transcription factors play an important role in everything from stress resistance to plant growth and development. However, the MYB family of chrysanthemums has not been the subject of a detailed bioinformatics and expression investigation. Results: In this study, we examined 324 CnMYB transcription factors from Chrysanthemum nankingense genome data, which contained 122 Cn1R-MYB, 183 CnR2R3-MYB, 12 Cn3R-MYB, 2 Cn4R-MYB, and 5 atypical CnMYB. The protein motifs and classification of CnMYB transcription factors were analyzed. Among them, motifs 1, 2, 3, and 4 were found to encode the MYB DNA-binding domain in R2R3-MYB proteins, while in other-MYB proteins, the motifs 1, 2, 3, 4, 5, 6, 7, and 8 encode the MYB DNA-binding domain. Among all CnMYBs, 44 genes were selected due to the presence of CpG islands, while methylation is detected in three genes, including CnMYB9, CnMYB152, and CnMYB219. We analyzed the expression levels of each CnMYB gene in ray floret, disc floret, flower bud, leaf, stem, and root tissues. Based on phylogenetic analysis and gene expression analysis, three genes appeared likely to control cellulose and lignin synthesis in stem tissue, and 16 genes appeared likely to regulate flowering time, anther, pollen development, and flower color. Fifty-one candidate genes that may be involved in stress response were identified through phylogenetic, stress-responseve motif of promoter, and qRT-PCR analyses. According to genes expression levels under stress conditions, six CnMYB genes (CnMYB9, CnMYB172, CnMYB186, CnMYB199, CnMYB219, and CnMYB152) were identified as key stress-responsive genes. Conclusions: This research provides useful information for further functional analysis of the CnMYB gene family in chrysanthemums, as well as offers candidate genes for further study of cellulose and lignin synthesis, flowering traits, salt and drought stress mechanism.
... In soybean, histone acetylation is an epigenetic mark involved in abiotic stress response ( Figure 3B). Song et al. (2012) demonstrated by ChIP analysis, that the activation of transcription factors responsive to salt, such as genes from MYB, b-ZIP, and AP2/DREB families was correlated with an increased level of histone H3K9 acetylation . ...
Article
Full-text available
Abiotic stresses such as nutritional imbalance, salt, light intensity, and high and low temperatures negatively affect plant growth and development. Through the course of evolution, plants developed multiple mechanisms to cope with environmental variations, such as physiological, morphological, and molecular adaptations. Epigenetic regulation, transcription factor activity, and post-transcriptional regulation operated by RNA molecules are mechanisms associated with gene expression regulation under stress. Epigenetic regulation, including histone and DNA covalent modifications, triggers chromatin remodeling and changes the accessibility of transcription machinery leading to alterations in gene activity and plant homeostasis responses. Soybean is a legume widely produced and whose productivity is deeply affected by abiotic stresses. Many studies explored how soybean faces stress to identify key elements and improve productivity through breeding and genetic engineering. This review summarizes recent progress in soybean gene expression regulation through epigenetic modifications and circRNAs pathways, and points out the knowledge gaps that are important to study by the scientific community. It focuses on epigenetic factors participating in soybean abiotic stress responses, and chromatin modifications in response to stressful environments and draws attention to the regulatory potential of circular RNA in post-transcriptional processing.
Article
Aluminum (Al) is the most abundant element in the earth crust. Due to the abuse of phosphate fertilizer, acid rain has been frequently observed in recent years, resulting in the conversion of non-toxic aluminosilicates in the soil into Al ions, thereby causing stress to plants. As a DNA methylation inhibitor, curcumin can effectively counteract the Al stress on plants, while the epigenetic mechanism remains unclear. This study discusses the epigenetic mechanism of curcumin counteracting Al stress on grape. The results demonstrated that curcumin could significantly relieve the Al stress symptoms of grapes and reduce its whole genome methylation level. Al stress and curcumin treatment did not cause variations in the methylation level in each chromosome. While Al stress led to a slight increase in the average methylation level of each chromosome, and treatment by curcumin led to a significant decrease in the average methylation level of each chromosome. Specifically, the sites of CG and CHG were decreased significantly, and the site of CHH was increased or decreased significantly. Analysis of differentially-methylated regions (DMRs) revealed that treatment by curcumin led to an increase in hypo-DMRs in the whole genome of grape, and analysis of differentially-methylated genes (DMGs) also identified differentially related genes of hypo-DMRs in the whole genome of grapes, suggesting that curcumin triggers responses to Al stress by regulating hypo-methylation mode of the whole genome of grape. Additionally, Kyoto Encyclopedia of Genes and Genomes (KEGG) enrichment analysis demonstrated that DMGs of grapes generate responses to Al stress by participating in galactose metabolism, ascorbate, and aldarate metabolism, and amino sugar and nucleotide sugar metabolism pathways of carbohydrate metabolism in the KEGG subclass.
Article
Full-text available
Abstracts Short-term ambient low temperature (ALT) stimulation is necessary for Osmanthus fragrans to facilitate continued flower opening after floral bud development reaches maturity. DNA methylation, a vital epigenetic modification, regulates various biological processes in response to temperature fluctuations. However, its role in temperature-driven flower opening remains elusive. In this study, we identified the pivotal timeframe during which O. fragrans promptly detected temperature cues. Using whole-genome bisulfite sequencing (WGBS), we explored global DNA hypomethylation during this phase, with the most significant changes occurring in CHH sequence contexts. Auxin transport inhibitor (TIBA) application revealed that ALT-induced endogenous auxin accumulation promoted peduncle elongation. In our mRNA-seq analysis, we discovered the differentially expressed genes (DEGs) with hypo-differentially methylated regions (hypo-DMRs) were mainly enriched in auxin and temperature response, RNA processing, carbohydrate and lipid metabolism. The transcripts of three DNA demethylase genes (OfROS1a, OfDML3, OfDME) showed upregulation. Furthermore, all DNA methylase genes, except for OfCMT2b, also displayed increased expression, specifically with two of them, OfCMT3a and OfCMT1, being associated with hypo-DMRs. Promoter assays showed that OfROS1a, with promoters containing low temperature and auxin responsive elements, were activated by ALT and exogenous IAA at low concentrations but inhibited at high concentrations. Overexpression OfROS1 reduced endogenous auxin levels but enhanced the expression of genes related to auxin response and spliceosome in petunia. Furthermore, OfROS1 promoted sucrose synthesis in petunia corollas. Our data characterized the rapid response of active DNA hypomethylation to ALT and suggested a possible epiregulation of temperature-dependent flower opening in O. fragrans.
Article
Full-text available
Plants are able to acclimate to new growth conditions on a relatively short time-scale. Recently, we showed that the progeny of plants exposed to various abiotic stresses exhibited changes in genome stability, methylation patterns and stress tolerance. Here, we performed a more detailed analysis of methylation patterns in the progeny of Arabidopsis thaliana (Arabidopsis) plants exposed to 25 and 75 mM sodium chloride. We found that the majority of gene promoters exhibiting changes in methylation were hypermethylated, and this group was overrepresented by regulators of the chromatin structure. The analysis of DNA methylation at gene bodies showed that hypermethylation in the progeny of stressed plants was primarily due to changes in the 5' and 3' ends as well as in exons rather than introns. All but one hypermethylated gene tested had lower gene expression. The analysis of histone modifications in the promoters and coding sequences showed that hypermethylation and lower gene expression correlated with the enrichment of H3K9me2 and depletion of H3K9ac histones. Thus, our work demonstrated a high degree of correlation between changes in DNA methylation, histone modifications and gene expression in the progeny of salt-stressed plants.
Article
Full-text available
Histone modifications play an important role in the epigenetic regulation of gene expression. All histone modifications are reversible, which may therefore provide a flexible way for regulating gene expression during the plant's development and during the plant response to environmental stimuli. The reversible acetylation and deacetylation of specific lysine residues on core histones are catalysed by histone acetyltransferases and histone deacetylases (HDAs). HDA6 is an RPD3-type histone deacetylase in Arabidopsis. The Arabidopsis HDA6 mutant, axe1-5, and HDA6 RNA-interfering plants displayed a phenotype that was hypersensitive to ABA and salt stress. Compared with wild-type plants, the expression of the ABA and abiotic stress-responsive genes, ABI1, ABI2, KAT1, KAT2, DREB2A, RD29A, and RD29B, was decreased in axe1-5 and HDA6 RNA-interfering plants when treated with ABA or salt stress. It was found that both ABA and salt stress could enrich the gene activation markers, histone H3K9K14 acetylation, and H3K4 trimethylation, but decrease the gene repression marker, H3K9 dimethylation, of the ABA and abiotic stress-responsive genes. Our study indicates that HDA6-involved histone modifications modulate seed germination and the salt stress response, as well as ABA- and salt stress-induced gene expression in Arabidopsis.
Article
In Arabidopsis, the induction of a dehydration-responsive gene, rd22, is mediated by abscisic acid (ABA). We reported previously that MYC and MYB recognition sites in the rd22 promoter region function as cis-acting elements in the drought- and ABA-induced gene expression of rd22. bHLH- and MYB-related transcription factors, rd22BP1 (renamed AtMYC2) and AtMYB2, interact specifically with the MYC and MYB recognition sites, respectively, in vitro and activate the transcription of the β-glucuronidase reporter gene driven by the MYC and MYB recognition sites in Arabidopsis leaf protoplasts. Here, we show that transgenic plants overexpressing AtMYC2 and/or AtMYB2 cDNAs have higher sensitivity to ABA. The ABA-induced gene expression of rd22 and AtADH1 was enhanced in these transgenic plants. Microarray analysis of the transgenic plants overexpressing both AtMYC2 and AtMYB2 cDNAs revealed that several ABA-inducible genes also are upregulated in the transgenic plants. By contrast, a Ds insertion mutant of the AtMYC2 gene was less sensitive to ABA and showed significantly decreased ABA-induced gene expression of rd22 and AtADH1. These results indicate that both AtMYC2 and AtMYB2 proteins function as transcriptional activators in ABA-inducible gene expression under drought stress in plants.
Article
We propose a model for heterochromatin assembly that links DNA methylation with histone methylation and DNA replication. The hypomethylated Arabidopsis mutants ddm1 and met1 were used to investigate the relationship between DNA methylation and chromatin organization. Both mutants show a reduction of heterochromatin due to dispersion of pericentromeric low-copy sequences away from heterochromatic chromocenters. DDM1 and MET1 control heterochromatin assembly at chromocenters by their influence on DNA maintenance (CpG) methylation and subsequent methylation of histone H3 lysine 9. In addition, DDM1 is required for deacetylation of histone H4 lysine 16. Analysis of F1 hybrids between wild-type and hypomethylated mutants revealed that DNA methylation is epigenetically inherited and represents the genomic imprint that is required to maintain pericentromeric heterochromatin.
Article
Mitogen-activated protein kinase (MAPK) cascades are known to transduce plant defense signals, but the downstream components of the MAPK have as yet not been elucidated. Here, we report an MAPK from rice (Oryza sativa), BWMK1, and a transcription factor, OsEREBP1, phosphorylated by the kinase. The MAPK carries a TDY phosphorylation motif instead of the more common TEY motif in its kinase domain and has an unusually extended C-terminal domain that is essential to its kinase activity and translocation to the nucleus. The MAPK phosphorylates OsEREBP1 that binds to the GCC box element (AGCCGCC) of the several basic pathogenesis-related gene promoters, which in turn enhances DNA-binding activity of the factor to the cis element in vitro. Transient co-expression of the BWMK1 and OsEREBP1 in Arabidopsis protoplasts elevates the expression of the β-glucuronidase reporter gene driven by the GCC box element. Furthermore, transgenic tobacco (Nicotiana tabacum) plants overexpressing BWMK1 expressed many pathogenesis-related genes at higher levels than wild-type plants with an enhanced resistance to pathogens. These findings suggest that MAPKs contribute to plant defense signal transduction by phosphorylating one or more transcription factors.
Article
Background: The heterochromatin of many eukaryotes is marked by both DNA methylation and histone H3 lysine 9 (H3-K9) methylation, though the exact relationship between these epigenetic modifications is unknown. In Neurospora, H3-K9 methylation is required for the maintenance of all known DNA methylation. In Arabidopsis, H3-K9 methylation directs some of the CpNpG and asymmetric methylation. However, it is not known in any organism whether DNA methylation may also direct histone H3 methylation.Results: Using chromatin immunoprecipitation (ChIP) assays, we show that Arabidopsis heterochromatin is associated with H3-K9 methylation. This histone methylation is dependent on the KRYPTONITE and DDM1 genes (SU[VAR]3-9 and SWI2/SNF2 homologs, respectively). We also find that a decrease in DNA methylation does not directly cause a loss of H3-K9 methylation. Instead, a decrease in H3-K9 methylation is only seen at loci where transcription is derepressed.Conclusions: We conclude that DNA methylation does not control the methylation of histone H3-K9. We propose that loss of H3-K9 methylation is due to transcriptional reactivation, coupled with deposition of unmethylated nucleosomes. These findings are consistent with recent observations of DNA replication-independent deposition of histone H3.3 in Drosophila. Our results also suggest that, in Arabidopsis, DNA methylation is sufficient for gene silencing, but H3-K9 methylation is not.
Article
The two most commonly used methods to analyze data from real-time, quantitative PCR experiments are absolute quantification and relative quantification. Absolute quantification determines the input copy number, usually by relating the PCR signal to a standard curve. Relative quantification relates the PCR signal of the target transcript in a treatment group to that of another sample such as an untreated control. The 2(-DeltaDeltaCr) method is a convenient way to analyze the relative changes in gene expression from real-time quantitative PCR experiments. The purpose of this report is to present the derivation, assumptions, and applications of the 2(-DeltaDeltaCr) method. In addition, we present the derivation and applications of two variations of the 2(-DeltaDeltaCr) method that may be useful in the analysis of real-time, quantitative PCR data. (C) 2001 Elsevier science.
Article
SummaryNAC transcription factors play important roles in plant growth, development and stress responses. Previously, we identified multiple NAC genes in soybean (Glycine max). Here, we identify the roles of two genes, GmNAC11 and GmNAC20, in stress responses and other processes. The two genes were differentially induced by multiple abiotic stresses and plant hormones, and their transcripts were abundant in roots and cotyledons. Both genes encoded proteins that localized to the nucleus and bound to the core DNA sequence CGT[G/A]. In the protoplast assay system, GmNAC11 acts as a transcriptional activator, whereas GmNAC20 functions as a mild repressor; however, the C-terminal end of GmANC20 has transcriptional activation activity. Over-expression of GmNAC20 enhances salt and freezing tolerance in transgenic Arabidopsis plants; however, GmNAC11 over-expression only improves salt tolerance. Over-expression of GmNAC20 also promotes lateral root formation. GmNAC20 may regulate stress tolerance through activation of the DREB/CBF–COR pathway, and may control lateral root development by altering auxin signaling-related genes. GmNAC11 probably regulates DREB1A and other stress-related genes. The roles of the two GmNAC genes in stress tolerance were further analyzed in soybean transgenic hairy roots. These results provide a basis for genetic manipulation to improve the agronomic traits of important crops.